<<

arXiv:1709.04428v1 [math.NT] 13 Sep 2017 oe h rgnlvrino h rbe a u oEwr Wa Edward to due was problem the Algebraicae of prob version central original a li The be extensive to come. an continue to with likely history seems It long . very a has problem Waring’s h u of sum the o esr xcl htWrn a nmn,w sueta em he that assume we mind, in had Waring what exactly sure be not ibr rvdteeitneof existence the proved Hilbert eaentrlt s hte h bound the whether t ask in to developed natural theory became number additive and integers Gaussian fteefrso aymteaiin vrapro fmn ye many of period a over mathematicians many of efforts the of many er ale neape ppaigi i treatise his in apppearing examples in earlier Dio years to known already by apparently was proven but was conjecture, that general theorem four-squares the to equivalent is enfudfralbtfiieymn ausof values many finitely but all for found been mletpstv integer positive smallest iie nsl ocnieigol those only considering to oneself limited novsfidn pe onsfor bounds upper finding involves mrvdbud n h usino nigepii auso values explicit finding of question the and bound, improved h rbe aigbe ovdol for only solved been having problem the hswr a atal upre yNAgatH98230-15-1- grant NSA by supported partially was work This k th M 00SbetClassification: Subject 2010 AMS Keywords: theo graph fields. spectral finite using for achieved are that fields finite lemma over Hensel’s and theory Artin-Wedderburn of combination A oeso aua ubr,satisfying numbers, natural of powers BSTRACT 9 ulse in published , ue;eeynme stesmof sum the is number every cubes; nti ae eoti hr eut o aigspolmo problem Waring’s for results sharp obtain we paper this In . aigsPolm pcrlGahTer,Artin-Wedderbur Theory, Graph Spectral Problem, Waring’s g ( 1770 k ) ihtepoet htevery that property the with , AIGSPOLMI IIERINGS FINITE IN PROBLEM WARING’S g hr esae Eeynme stesmof sum the is number "Every states he where , ( k G ) o arbitrary for ( YE¸S rmr:1P5 eodr:0C5 16K99. 05C25, Secondary: 11P05; Primary: k ) k sn ainso h ad-iteodcrl ehd whic method, circle Hardy-Littlewood the of variants using MDEM IM ˙ x g 2 = ( htaesfcetylre elet We large. sufficiently are that .I 1. k ) k and ol eipoe fisedo osdrn all considering of instead if improved be could g NTRODUCTION 19 se[8) ihmlil ra fmteaissc sthe as such mathematics of areas multiple with [18]), (see IRO ˙ 2 4 = (2) k Arithmetica k iudae [ biquadrates yeautn a evaluating by L KARABULUT GLU ˘ 1 4 = 0319. y eas rv naaou fSröystheorem Sárközy’s of analogue an prove also We ry. , n fteapoce o takn hsquestion this attacking for approaches the of One . g 3 9 = (3) n uligo e roso nlgu results analogous of proofs new on building and x hnu h lue oi eryfite hundred fifteen nearly it to alluded who phantus e nmteaia eerhfrsm ieto time some for research mathematical in lem arnetesm erWrn ulse his published Waring year same the Lagrange epoeso bann hs eut.I then It results. these obtaining of process he f ∈ se[15]). (see G 4 eauesann aydfeetaesof areas different many spanning terature N ( th r,epii omlsfor formulas explicit ars, k a ewitna u fa most at of sum a as written be can ig n rtapae in appeared first and , , ) atta o every for that eant hoy iieRings. Finite Theory, n 25 oes;ads n" lhuhw can we Although on.". so and powers]; g o arbitrary for 4 19 = (4) e eea nt ig,b sn a using by rings, finite general ver fl nerlin integral -fold G ( t.Tefc that fact The etc. , k 4 ) qae;eeynme is number every squares; k eoetesals such smallest the denote ssilwd pn with open, wide still is 1909 k > 2 n saresult a as and , g hr xssa exists there , ( Meditationes k x ) aenow have ∈ g 2 4 = (2) N sa is h one , g ( k ) 2 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT powerful analytic tool also motivated by Waring’s problem (for an overview of the complete history of the problem, see [18]).

After considering this question in the context of N, it seems natural to generalize it to working over Zn, traditionally referred to as Waring’s problem mod n. One can fix a k and look for a result which gives the th smallest m = g(k,n) such that every number in Zn can be written as a sum of at most m many k powers. This problem was first completely solved for k = 2, 3 by Small in [16], after which he observed that the same techniques yield a solution for all k, [13]. 1

The main contribution of the current paper is to extend results for Waring’s problem to the context of general finite rings. As far as we are aware, there do not exist other such results in the literature, so that our results appear to be the first known ones of this type. We get these generalizations by appealing to Artin-Wedderburn theory and Hensel’s lemma, allowing us to build on results over finite fields obtained at an earlier point (see Section 3) in our paper.

In particular, the following represents a sampling of the type of results found in Section 5, (see Theorem 5.2 for a more complete statement).

Let R be a (not necessarily commutative) finite ring. Then any element of R can be written as a sum of n many kth powers in R

where n = 2, if k = 3 as long as 3, 4, 7 ∤ R • | | where n = 5, if k = 4 as long as 2, 9 ∤ R • | | where n = 3, if k = 5 as long as 5, 11, 16 ∤ R • | | where n = 7, if k = 6 as long as 2, 3, 25 ∤ R • | | where n = 4, if k = 7 as long as 7, 8 ∤ R • | | where one can lower the values of n for any given k as long as one is willing to exclude a few more numbers from dividing R . | |

In order to save space, we limit our explicit statements in Theorem 5.2 to 3 6 k 6 11, but similar results could easily be obtained for any value of k for which one has results over finite fields. This reduces the question of Waring’s problem over general finite rings to the analogous questions over finite fields.

Note that Small’s solution to Waring’s problem mod n when n is a prime answers Waring’s problem for a special class of finite fields. In addition, Small obtains a result for more general finite fields by applying an inequality from the theory of diagonal equations over such fields to prove that if q > (k 1)4, then − 1There are a few minor errors in the stated values of g(k, n) that appear in [13], [14], [16]. The correct values should be g(4, 41) = 2, g(5, 71) = 2 and g(5, 101) = 2. WARING’S PROBLEM IN FINITE RINGS 3

th every element of Fq can be written as a sum of two k powers in [14] (which when translated into the language of this paper says γ(k,q) 6 2 whenever q > (k 1)4). In Section 3, we deduce an asymptotically − equivalent result to his and to results of [19], [21] by applying graph theoretic methods yielding self contained elementary proofs that when q > k4, every element of the field can be written as a sum of two kth powers, and when q > k3, as three kth powers, etc. We also enlist the help of a supercomputer to calculate the values for γ(k,q) for all q too small to be covered by the above mentioned results when k 6 37.

On top of this, in the process of proving our finite field Waring’s problem results (similar to results that were already known), we also obtain the following new result providing an analogue of Sárközy’s theorem in the finite field setting. qk Theorem. Let k be a positive integer. If E is a subset of F with size E > then it contains at least q | | √q 1 th − two distinct elements whose difference is a k power. Thus, in particular if k is fixed and E = Ω √q as | | q , then E contains at least two distinct elements whose difference is a kth power. →∞ 

Our main tool in the proofs of our finite field and Sárközy’s results is a spectral gap theorem for Cayley digraphs for which we provide a self-contained proof in Section 2.1. By building on these finite field results, we also obtain in Section 4 upper bounds for Waring numbers for matrix rings over finite fields (hence over general semisimple finite rings) that are stronger than our aforementioned results obtained in the case of more general finite rings. Results relating to matrix rings exist in the literature (see the history section and references in [10] for a summary), but they often focus on existence results, showing that a given matrix over a can be written as a sum of some number of kth powers, rather than determining how many kth powers need to be used. As was the case for our more general results mentioned above, our results over matrix and semisimple finite rings also rely on an efficient use of Hensel’s Lemma, (see Theorem 4.2 for more details).

2. PRELIMINARIES

2.1. Spectral Graph Theory. The material presented here prior to Cayley digraphs is pretty standard and can be found in any introductory graph theory book, e.g. [7], [20].

A graph G consists of a vertex set V (G), an edge set E(G), and a relation that associates with each edge two vertices called its endpoints.A directed graph or digraph is a graph in which the endpoints of each edge is an ordered pair of vertices. If (u, v) also denoted by u v is such an ordered pair of vertices, that means → there exists an edge in the digraph with tail u and head v. When we draw a digraph, we represent the vertices with some dots and the directed edges between the vertices with some curves on the plane, and we give each edge a direction from its tail to its head. We define the out-degree of a vertex u, denoted by d+(u), as the number of edges with tail u and the in-degree, d−(u), as the number of edges with head u. Each loop (i.e. 4 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT an edge have the same endpoints) contributes +1 to both in- and out-degree. As a side note, notice that the summation of in-degrees and out-degrees are always equal to each other, and equals to the number of edges.

We define the adjacency matrix AG (or A) of a digraph G as a matrix with rows and columns indexed by the vertices of G, such that the uv-entry of A is equal to the number of edges with tail u and head v. Since AG is not necessarily symmetric, unfortunately we lose the spectral theorem of linear algebra in most cases when

G is a digraph, but we can still consider the spectrum of AG and it turns out to be still very useful.

If G is a digraph, a walk of length r in G can be defined as a sequence of vertices (v , v , , v ) with the 0 1 · · · r only restriction that the ordered pair (vi 1, vi) is the endpoints of some edge for every 1 6 i 6 r. That is − to say if we want to walk from vi 1 to vi, we need the existence of an edge with the tail vi 1 and head vi, − − i.e. the direction matters in digraphs in contrast to graphs. On the other hand the definition of walk allows us to go over any edge and any vertex more than once. By induction on k we can show that the uv-entry of Ak, where A denotes the adjacency matrix of G, counts the (directed) walks of length k from u to v. It is also easy to show that if G is a simple digraph, i.e. a digraph with no multi-edges and with at most one loop, T + then the trace of AA equals u G d (u)= E(G) . Here multi-edge means there exists at least two edges ∈ | | which correspond to the same endpoints u v. P →

Let H be a finite and S be a subset of H. The Cayley digraph Cay(H,S) is the simple digraph whose vertex set is H and u v if and only if v u S. By definition Cay(H,S) is a simple digraph → − ∈ with d+(u)= d (u)= S . Furthermore, if we have a Cayley digraph, then we can find its spectrum easily − | | using characters from representation theory, see [12] for a rigorous treatment of character theory. A function χ : H C is a character of H if χ is a group homomorphism from H into the C . −→ ∗ If χ(h) = 1 for every h H, we say χ is the trivial character. The following theorem is a very important ∈ well-known fact, see e.g. [4]:

Theorem 2.1. Let A be an adjacency matrix of a Cayley digraph Cay(H,S). Let χ be a character on H.

Then the vector (χ(h))h H is an eigenvector of A, with eigenvalue s S χ(s). In particular, the trivial ∈ ∈ character corresponds to the trivial eigenvector 1 with eigenvalue S . | |P

Proof. Let u , u , , u be an ordering of the vertices of the digraph and let A correspond to this ordering. 1 2 · · · n Pick any ui. Then we have

n Aijχ(uj)= χ(uj)= χ(ui + s)= χ(ui)χ(s)= χ(s) χ(ui). j=1 ui uj s S s S "s S # X X→ X∈ X∈ X∈ Thus

Aχ = χ(s) χ. "s S # X∈  WARING’S PROBLEM IN FINITE RINGS 5

Notice that we get H many eigenvectors from the characters as demonstrated in the previous theorem, | | and they are all distinct since they are orthogonal by character orthogonality, [12]. This means adjacency matrix A of a Cayley digraph is diagonalizable and we know all of the eigenvectors explicitly, assuming we know all of the characters of H. This shows us that A is non-defective and the eigenvectors of A form an eigenbasis. Indeed, it will be very important in the proof of the spectral gap theorem (upcoming). Recall when we have a (undirected) graph, since we have A real and symmetric, we know there exists an eigenbasis of A by the spectral theorem of linear algebra. On the other hand, this is not the case for digraphs in general. For example, consider the digraph with vertex set V (G) = u, v and with only one directed edge u v. 0{ 1 } → In this case G is a simple digraph with no loops; A = has only one eigenvector, so A does not G 0 0 G have an eigenbasis. This makes Cayley digraphs nicer than general simple digraphs.

Theorem 2.2 (viz. spectral gap theorem) below is a very important and widely used tool in graph theory by itself. Hence we intend to give a self-contained proof.

Theorem 2.2 (Spectral Gap Theorem For Cayley Digraphs). Let Cay(H,S) be a Cayley digraph of order n. Let χi i=1,2, ,n be the set of all distinct characters on H such that χ1 is the trivial one. Define { } ··· n n = max χi(s) ∗ S 26i6n ! | | s S X∈ and let X,Y be subsets of vertices of Cay(H,S). If X Y > n , then there exists a directed edge | || | ∗ between a vertex in X and a vertex in Y . In particular if X > n , then there exists at least two distinct p | | ∗ vertices of X with a directed edge between them.

Proof. Let u , u , , u be an ordering of the vertices of Cay(H,S) and let A be the adjacency matrix 1 2 · · · n corresponding to this ordering. By Theorem 2.1 each χi gives us an eigenvector for A with an eigenvalue

s S χi(s). We can normalize χi for each i 1, 2, ,n and get vi’s so that v1, , vn is an orthonor- ∈ ∈ { · · · } · · · mal basis for Cn. Therefore, any complex n-dimensional vector v can be written as v = n v, v v P j=1h j i j where , denotes the Hermitian inner product in Cn. Notice h− −i P n n n 2 v, v = v, vj vj, v = v, vj vj , v = v, vj (1) h i * h i + h ih i kh ik Xj=1 Xj=1 Xj=1 which is also known as the Plancherel identity.

Define 1X as the 0-1 column vector whose ith entry is 1 when u X and 0 otherwise. Define 1Y similarly. i ∈ If we calculate

a1 a2 A1Y =  .  .   an     6 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT then each ai denotes the number of directed edges from ui to the vertices of Y . If we multiply both sides of T the equation with 1X from the left, we get

a1 T T a2 1X A1Y = 1X  .  = ai . ui X   X∈ an   which is indeed exactly the number of edges from the vertices of X to the vertices of Y . This calculation T shows that as long as 1X A1Y > 0 there will be a directed edge from a vertex in X to a vertex in Y .

On the other hand we also have

n n T 1X A1Y = 1X , A1Y = 1X , A 1Y , vj vj = 1X , 1Y , vj λjvj h i *  h i + *  h i + Xj=1 Xj=1 n  n  n  = 1X , 1Y , v λ v = 1X , λ v 1Y , v = λ 1X , v v , 1Y (2) h h ji j ji h j jih ji jh jih j i Xj=1 Xj=1 Xj=1

By Theorem 2.1 we know one of the eigenvalues of Cay(H,S) is λ = S with eigenvector v = 1 1 and 1 | | 1 √n by substitution in (2) we get n T X Y S 1X A1Y = | || || | + λ 1X , v v , 1Y . n jh jih j i Xj=2 Let n E = λ v , 1Y 1X , v . jh j ih ji Xj=2 X Y S We will show if X Y >n then | || || | > E , and the result will follow. | || | ∗ n | | p By the Cauchy-Schwarz inequality we have

1/2 1/2 n n 2 2 E 6 ( max λj ) 1Y , vj 1X , vj | | 26j6n k k  kh ik   kh ik  Xj=1 Xj=1     Moreover by Plancherel’s equality we have

E 6 max λj X Y . | | 26j6n k k | || |   p Hence, the result follows as long as S X Y | || || | > max λj X Y n 26j6n k k | || |   p holds. The only thing left to point out is that in the statement of the theorem we substituted λj’s using Theorem 2.1.  WARING’S PROBLEM IN FINITE RINGS 7

Keep the notation in the previous theorem for the next corollary. Define k n n ,k = k max χi(s) . ∗ S 26i6n ! | | s S X∈

Corollary 2.3. If n ,k < X Y , then there exists a directed k-walk from a vertex in X to a vertex in ∗ | || | Y . In particular if n ,k < X , then there exists at least two distinct vertices of X with a directed k-walk ∗ p| | between them.

The proof of the corollary follows from a well-known trick in graph theory. We form a new graph with the same vertex set Cay(H,S), where every directed edge from u to v in the new graph corresponds to an oriented walk of length k from u to v in the original digraph. This new graph can be most easily defined as the digraph with the same vertex set as new graph and adjacency matrix Ak where A is adjacency matrix of the original graph. Notice Ak has the same orthogonal eigenbasis A does and its eigenvalues are the kth powers of A’s. So, the corollary follows when we apply the theorem on the new graph. Furthermore, notice that we can prove this theorem for any digraph with orthogonal eigenbasis, and not just for Cayley digraphs.

Corollary 2.4 (Spectral Gap Theorem For d-regular Graphs). Let G be a d-regular simple graph with eigen- values λ > > λ . Let 1 · · · n n n = max λi ∗ d 26i6n | |   and let X,Y V (G) be subsets of vertices. If X Y >n , then there exists an edge incident to a vertex ⊂ | || | ∗ in X and a vertex in Y . In particular if X > n , then there exists at least two vertices of X adjacent to | | p ∗ each other.

This corollary does not follow from the statement of the theorem directly, but it follows from the proof. There are only a few points to change in the theorem’s proof to prove the corollory. First, we do not always know the eigenvalues and eigenvectors of a regular simple graph explicitly. But we do know if G is a d-regular simple graph with n vertices, then it should have λ = d as an eigenvalue with multiplicity at least 1 and n T there exists an eigenbasis for R . Second, since the 1X A1Y term in the theorem’s proof was the number of directed edges; it will be the number of edges between the vertices of X and vertices of Y such that the edges in the intersection counted twice, where the edges in the intersection refers to the edges which has the endpoints in X Y . However, the rest of the proof works as demonstrated above. ∩

2.1.1. Cayley Graphs and Character Theory. We need to use characters only for finite fields in this paper.

Let Fp be the finite field of prime order p. Every time we fix a j 0, 1, ,p 1 and define χj(x) := 2πijx ∈ { · · · − } e p for every x F , χ becomes a character on the additive group structure of F , [9]. Since different j ∈ p j p values provide us different characters on Fp, we have p many characters defined in this way, but the theory 8 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT tells us we should have p many characters for a group of order p, so we obtained all the characters on

Fp.

Let Fq be the finite field of order q with p and Fq∗ denote the multiplicative group of it as usual.

We want to give a characterization for the characters on Fq. Once we know the characters for the base field F , the characters of F can be obtained by following the process we are about to explain. Let α F and p q ∈ q∗ define a map φ : F F as φ (x) := tr(αx) for every x F . Here αx denotes the multiplication of α α q −→ p α ∈ q and x in F , and tr is the usual field trace of F over F . We can easily show that for any element α F and q q p ∈ q∗ for any character χ on F , χ φ defines a character on F . We can also show that χ φ is the trivial j p j ◦ α q j ◦ α character on F if and only if j is zero. Then also notice that χ φ and χ φ is the same character if and q j ◦ α j ◦ β only if α = β. Hence, if we fix j = 1 and if we range α in F , χ φ ’s become distinct. That is why we q∗ 1 ◦ α have q 1 many distinct nontrivial characters by ranging α. We also have the trivial character, and we can − think of the trivial character as j = 1 and α = 0. In this way we obtain at least q many distinct characters on Fq, but from the theory we know we should have exactly q many characters on Fq, which means any character on F can be written explicitly as χ φ for some α F . q 1 ◦ α ∈ q

2.2. Finite Rings, Jacobson Radical and Artin-Wedderburn Theory. As we stated before, the main purpose of this paper is to solve Waring’s problem for finite rings with identity. We actually have a wide range of such rings. Some important examples are finite fields, the ring of integers modulo n for each integer n > 2, matrix rings over finite fields, which we will denote by Fq, Zn, Matn(Fq), respectively. Some subrings of Matn(Fq) (e.g. upper triangular matrices, lower triangular matrices, diagonal matrices etc., or F [A] for A Mat (F )) are also familiar examples. A particular class of examples, which will be used in q ∈ n q Section 4, comes from quotients of polynomial rings over finite fields with some principal ideal generated by F a polynomial over the same field i.e. q[x] f(x) . If the polynomial is irreducible over Fq, the quotient h i ring is isomorphic to a finite field. Otherwise. it is not even an integral domain. In addition to those when we take a finite family of finite rings with identity, we can make the direct sum of underlying abelian groups into a ring by defining multiplication coordinate-wise and the resulting ring is a finite ring with identity.

Another important family of examples for finite rings with identity come from group rings. When R is a ring with identity and G is a group with operation , the group ring R[G] is the set of all linear combinations ∗ of finitely many elements of G with coefficients in R i.e. any x R[G] can be written as g G agg ∈ ∈ where it is assumed that ag = 0 for all but finitely many elements of G. The addition is defined byP the rule g G agg+ g G bgg = g G(ag +bg)g and the multiplication is given by g G agg g G bgg = ∈ ∈ ∈ ∈ ∈ g G, h H (agbh)g h. The group ring of a finite group over any finite ring (or finite field)  is a finite ring. P ∈ ∈ P ∗ P P P WhenP R is a field, R[G] is called a group algebra. When the characteristic of the finite field divides the order of the group, the group algebra breaks up into finite rings called block rings.

Artin-Wedderburn theory provides us very useful structure theorems for certain classes of rings. We now recall the basics of this theory. The omitted proofs are standard and can be found in most (noncommutative) WARING’S PROBLEM IN FINITE RINGS 9 algebra books presenting the subject. In particular see [6], [8] since our treatment follows theirs closely for the convention and notation.

Let R be any ring, not necessarily finite. Then there is a left ideal J(R), or just J, called the Jacobson radical of R such that J = m where is the set of all maximal left ideals. Let x R. It is a well-known fact m M ∈ that x J if and onlyT∈M if 1+ rxs is a unit in R for all r, s R. We can also prove that J is the intersection of ∈ ∈ all maximal right ideals of R. This combined with the definition of J together implies that J is a two-sided ideal. Also notice that 1 / J so J = R. ∈ 6

A ring R is said to be semisimple if its Jacobson radical J is zero. As we state later, Artin-Wedderburn theory gives us Mat (F ) Mat (F ), the direct product of finitely many matrix rings over finite fields, n1 q1 ×···× nj qj is a semisimple finite ring. Maschke’s theorem shows many group algebras are semisimple, [5]:

Theorem 2.5 (Maschke’s Theorem). Let H be a finite group and let F be a field whose characteristic does not divide the order of H. Then F [H], the group algebra of H, is semisimple.

In the case when the characteristic of the finite field divides the order of the group, Maschke’s theorem does not apply and the group algebra is no longer semisimple. Instead it decomposes into block rings, which are another class of finite rings, with great importance in modular representation theory, see [1].

We now state the main theorem in Artin-Wedderburn theory that shows in particular that any finite semisim- ple ring is a product of matrix rings.

Theorem 2.6 (Artin-Wedderburn Theorem). The following conditions on a ring R are equivalent:

R is a nonzero semisimple left . • R is a direct product of a finite number of simple ideals each of which is isomorphic to the endomor- • phism ring of a finite dimensional over a .

There exist division rings D , , D and positive integers n , ,n such that R is isomorphic to • 1 · · · t 1 · · · t the ring Mat (D ) Mat (D ) Mat (D ). n1 1 × n2 2 ×···× nt t

While in the general theory of Artin-Wedderburn division rings appear, we need not work with division rings as a result of the following theorem.

Theorem 2.7 (Wedderburn’s Little Theorem). Every finite division ring D is a field.

We will also use Nakayama’s lemma:

Theorem 2.8 (Nakayama’s Lemma). If I1 is an ideal in a ring R with identity, then the following conditions are equivalent: 10 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

I is contained in the Jacobson radical, J • 1 If I is a finitely generated R-module such that I I = I , then I = 0. • 2 1 2 2 2 Corollary 2.9. If R is a finite ring, then J l = 0 for some l Z . ∈ +

Proof. Since R is finite, the following chain has to stop for some l Z ∈ + J J 2 J 3 J l = J l+1 = J l+2 = . ⊇ ⊇ ⊇···⊇ · · · l l If J = 0, we are done. Otherwise, let I1 = J, I2 = J and apply Nakayama’s lemma. 

Actually we do not need R to be a finite ring to have J l = 0 for some l, if R is just Artinian the result still follows by the same proof. We record this version since we need this result later only when R is finite.

2.3. Hensel’s Lemma in Special Cases. In this section we present and prove Hensel’s lemma for the Ja- cobson radical of finite commutative rings with identity and polynomial rings over finite fields. These results of course can be deduced from the general theory easily, but we provide self contained proofs. Our approach follows the standard proof of Hensel’s lemma in other settings, which usually resemble the classical Newton approximation method, e.g. see Hensel’s lemma for the p-adic integers in [11].

2.3.1. Hensel’s Lemma for the Jacobson Radical. Let R be a ring with identity and let J denote the Jacob- son radical of R. First, we record two general basic facts in the following lemmata, since we need them later.

i Lemma 2.10. If a + J is a left unit in R /J , then a is a left unit in R and this implies a + J is a left unit in R /J i for any i > 1.

Proof. If a + J is a left unit in R /J , then there exists some u R such that (u + J)(a + J)=1+ J. That ∈ means there exists some u R and j J such that ua =1+ j. This implies that there exists some y R ∈ ∈ ∈ such that yua = 1, which says yu is a left inverse of a. Also notice that (yu + J i)(a + J i) := yua + J i =1+ J i which means a + J i is a left unit in R /J i . 

One can replace left with right or 2-sided, and both the statement and proof of the previous proposition extend mutatis mutandis. As a side note also recall that the existence of a left inverse for an element of R in general does not imply the existence of a right inverse, and vice versa. But finite rings are nicer in this regard, we can prove that if there exists a left (resp. right) inverse for r R, then the left (resp. right) inverse ∈ of r is also the right (resp. left) inverse of r.

Lemma 2.11 (Taylor’s expansion of a polynomial). Let R be a commutative ring and p(x) R[x] be any ∈ polynomial. There exists a polynomial p R[x,y] such that p(x + y)= p(x)+ yp (x)+ y2p (x,y). 2 ∈ ′ 2 WARING’S PROBLEM IN FINITE RINGS 11

n Proof. Let p(x)= anx . Then, n n n 1 2 n n 1 2 p(x + y)= a (Px + y) = a (x + nx − y + y ( )) = a x + y na x − + y p (x,y). n n · · · n n 2 X X X X 

Proposition 2.12. Let R be a commutative finite ring with identity. Let p(x) R[x] where R[x] stands for ∈ the polynomial ring over R. Assume there exists a root of p(x) in R /J , say a0 + J for some a0 R, i.e. ∈ p(a0) p(a0 + J) 0 mod J. If p′(a0)+ J is a unit in R /J , then r = a0 satisfies ≡ − p′(a0) (1) p(r) 0 (mod J 2) ≡

(2) r = a0 (mod J)

(3) p′(a0) is a unit in R and p′(r) is a unit in R.

1 Proof. Notice that the existence of follows from Lemma 2.10 and from the assumptions. Suppose we p′(a0) have a root of p(x) in R /J 2 , let’s denote it with r + J 2. If we send r + J 2 using the canonical reduction map from R /J 2 to R /J , the image of r + J 2 should also be a root i.e. p(r + J 2) 0 (mod J 2) = p(r + J) 0 (mod J). ≡ ⇒ ≡ 2 If r + J is a lift of a0 + J then we should have r + J = a0 + J in R /J i.e. r = a0 + j for some j J. ∈ By Lemma 2.11, we have 2 p(r)= p(a0 + j)= p(a0)+ p′(a0)j + j b for some b R. We need p(a + j) 0 (mod J 2) that means we need ∈ 0 ≡ 2 p(a )+ p′(a ).j 0 (mod J ). 0 0 ≡ p(a ) So, if we let j = 0 from the beginning, everything works out in these equations and the claims −p′(a0) follow. 

2 If a0 + J is a root of p(x) in R /J and p′(a0) is a unit in R, then we can upgrade a0 + J to a root r + J 2 in R /J 2 by the proposition. But notice that using the exact same process, we can lift r + J in R /J 2 to 3 some root r1 + J in R /J 3 and so on. This process gives us a solution in each R /J i for any i > 1, and combined with Corollary 2.9, we obtain the following corollaries.

Corollary 2.13. Let R be a finite commutative ring with identity and p(x) R[x]. If there exists a root of ∈ p(x), say a0 + J, in R /J and p′(a0)+ J is a unit in R /J , then there exists a root of p(x) in R.

k k Corollary 2.14. Let R be a finite commutative ring with identity. Let α R. If α B1 + + Bm mod J k 1 ∈ ≡ · · · in R /J for some B1,B2, ,Bm R and kB − is a unit in R, then there exists some β in R such that · · · ∈ 1 α = βk + Bk + + Bk holds. 2 · · · m

Proof. Let p(x)= xk + Bk + + Bk α and apply the previous corollary.  2 · · · m − 12 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

2.3.2. Hensel’s Lemma for Polynomial Rings. Let Q(t) be a polynomial in Fq[x][t]= Fq[x,t]. We can write Q(t) as a + a t + a t2 + + a tn for some a , a , a F [x]. This means we may evaluate Q(t) at 0 1 2 · · · n 0 1 · · · n ∈ q any t F [x] and get outputs in F [x]. ∈ q q

Let f(x) be an irreducible polynomial in Fq[x]. Let g be any polynomial in Fq[x]. We can define f-adic valuation of g, say ν(g), such as ν(g)= m g 0 (mod f m) and g 0 (mod f m+1) for some m Z . ⇐⇒ ≡ 6≡ ∈ + We also set ν(g) = 0 if f ∤ g.

Proposition 2.15. Let g F [x] satisfy Q(g) 0 (mod f n) for some n Z . Suppose m is a positive ∈ q ≡ ∈ + integer less than or equal to n. If Q (g) 0 (mod f), there exists a g F [x] such that Q(g ) 0 ′ 6≡ 2 ∈ q 2 ≡ (mod f n+m) and g g (mod f n). Moreover, g is unique modf n+m. ≡ 2 2

n+m We will refer g2 as a lifting of g in modf .

Remark. We should note that the restriction m 6 n in this proposition is not that important. Since if we want to find a solution modulo f j for some j>n, we can apply the proposition j n many times recursively for − m = 1 and lift a solution to modulo f j. The same idea also applies to the next proposition.

Proof. By setting Q(g + f nt) 0 (mod f n+m), we can solve for t. By applying Lemma 2.11 on Q we ≡ know that

n n n 2 n Q(g + f t)= Q(g)+ f tQ′(g) + (f t) Q˜(g,f t) for some polynomial Q˜ n n+m Q(g)+ f tQ′(g) (mod f ) since n > m. ≡ By assumption Q(g) 0 (mod f n) i.e. Q(g)= f nh for some h F [x]. So, we should have ≡ ∈ q n n n n+m Q(g + f t) f h + f tQ′(g) (mod f ) ≡ n n+m f (h + tQ′(g)) (mod f ). ≡ Moreover,

n n+m n n+m Q(g + f t) 0 (mod f ) f (h + tQ′(g)) = f v for some v F [x] ≡ ⇐⇒ ∈ q n m f (h + tQ′(g) f v) = 0 for some v F [x]. ⇐⇒ − ∈ q To have this, since F [x] is an integral domain, we need h + tQ (g) f mv to be zero, that means we need q ′ − m h m n h h + tQ′(g) = f v which is the same with t − (mod f ). So, if we define g2 = g f = ≡ Q′(g) − Q′(g) Q(g) g from the beginning, we are all set. − Q′(g)

The only point we did not justify here is why Q (g) 1 makes sense in modf m. Remember Q (g) F [x]. ′ − ′ ∈ q F [x] is a PID, that means we have Bézout’s identity in F [x]. By assumption we have Q (g) 0 (mod f), q q ′ 6≡ which means f doesn’t divide Q′(g). Since f is irreducible, the only divisors of f are 1 and itself. So, WARING’S PROBLEM IN FINITE RINGS 13

m m gcd(Q′(g),f) = 1 and this implies gcd(Q′(g),f ) = 1. By Bézout, we have AQ′(g)+ Bf = 1 for some A, B F [x] which shows Q (g) 1 exits in modf m for any m > 1.  ∈ q ′ −

n Proposition 2.16. Let g Fq[x] be such that Q(g) 0 (mod f ) for some n Z+, and let m denote ∈ Q(g) ≡ ∈ ν(Q′(g)). If n> 2m, then g2 = g satisfies − Q′(g) (1) Q(g ) 0 (mod f n+1) 2 ≡ (2) g g (mod f n m) 2 ≡ −

(3) ν(Q′(g2)) = ν(Q′(g)) = m.

Proof. Since Q(g) 0 (mod f n), there exists an h F [x] such that Q(g) = f nh . Similarly since ≡ 1 ∈ q 1 ν(Q (g)) = m, there exists an h F [x] such that Q (g)= f mh and h 0 (mod f), which implies h ′ 2 ∈ q ′ 2 2 6≡ 2 is a unit in modf i for any i.

To prove (1) by applying Lemma 2.11 on Q we have

Q(g) Q(g )= Q g 2 − Q (g)  ′  Q(g) Q(g) 2 = Q(g)+ − Q′(g)+ Qˆ( ) for some polynomial Qˆ Q (g) Q (g) · · ·  ′   ′  n m 1 2 = (f − h h− ) Qˆ( ) 1 2 · · · 0 (mod f n+1) since n> 2m. ≡

(2) follows from

n Q(g) f h1 n m 1 n m g2 g = − = − m = f − h1h2− 0 (mod f − ). − Q′(g) f h2 − ≡

To prove (3) by applying Lemma 2.11 on Q′ we have

Q′(g )= Q′ g + (g g) 2 2 − 2 = Q′(g) + (g g)Q′′(g) + (g g) Qˇ( ) for some polynomial Qˇ 2 − 2 − · · · Q(g) Q(g) 2 = Q′(g) Q′′(g)+ Qˇ( ) − Q (g) Q (g) · · ·  ′   ′  m n m 1 n m 1 2 = f h f − h h− Q′′(g) + (f − h h− ) Qˇ( ) 2 − 1 2 1 2 · · · m n 2m 1 2n 3m 1 2 = f h f − h h− Q′′(g)+ f − (h h− ) Qˇ( ) 2 − 1 2 1 2 · · · 0 (mod f m) since n> 2m.  ≡ n 2m 1 2n 3m 1 2 Since h f h h− Q (g)+ f (h h− ) Qˇ( ) 0 (mod f) we have ν Q (g ) = m.  2 − − 1 2 ′′ − 1 2 · · · 6≡ ′ 2    14 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

3. WARING’S PROBLEMIN FINITE FIELDS

Let F be the finite field with q = ps elements where p is the characteristic of F . We pick a k Z and q q ∈ + denote the set of all kth powers in F by R i.e. R = αk α F . Let γ(k,q) denote the smallest m q k k { | ∈ q} th such that every element of Fq can be written as a sum of m many k powers in Fq. We are interested in finding γ(k,q) for different k and q values, but unfortunately γ(k,q) does not always exist. For a fixed value th th of k, if every element of Fq can be written as a sum of k powers, we say Fq is coverable by the k powers. Otherwise, we call it an uncoverable field for that specific k.

Since F2 and F3 are coverable for any k, the smallest example of an uncoverable field is F4 with k = 3. The only cubes in F are 0, 1 and any linear combination of 0 and 1 is either 0 or 1, this shows if α is a 4 { } primitive cube root of unity in F4, then it cannot be written as a sum of cubes. Therefore, we say F4 is an uncoverable field for k = 3 or cubes. This example is a very simple observation, but it actually illustrates th the point very well which is if Rk lies in a proper subfield of Fq, then we cannot cover Fq with k powers. th Furthermore, the converse also holds meaning Fq is coverable if and only if the set of k powers does not lie in a proper subfield, see [21].

Before we proceed, notice that γ(1,q) is trivially 1. In the case of k = 2, if Fq is a field with characteristic 2, we have the Frobenius automorphism which implies every element of Fq is a square so that γ(2,q) = 1. If char F = 2, then exactly half of the elements in F are squares and γ(2,q) = 2 follows by the pigeonhole q 6 q∗ principle.

If we look at the k = 3 case, then one can prove that except for q = 4, 7 in every finite field every element can be written as a sum of at most two cubes, see [2]. F7 requires three cubes since the only cubes in F7 are 0, 1, +1 , and thus 3 is not the sum of any two cubes. F is uncoverable as we discussed above. { − } 4

We can also easily single out a nice bound for γ(k,q) when it exists, and also the extreme case i.e. when γ(k,q) = 1:

Proposition 3.1. Let Fq be the finite field with order q and let k be a positive integer. If Fq is coverable with the kth powers, then F dR where d = gcd(k,q 1). In particular if γ(k,q) exists, then it is less than or q ⊆ k − equal to k. Moreover, F = R if and only if gcd(k,q 1) = 1. q k −

Proof. We have R 2R mR = (m + 1)R = for some m since F is finite. We assumed k ⊆ k ⊆···⊆ k k · · · q F is coverable this implies mR = F . If nR = (n + 1)R for some n, that means there exists a nonzero q k q k 6 k element α in (n + 1)R but not in nR . Let R := R 0 . R is a group under multiplication and it is a k k k∗ k \ { } k∗ th subgroup of Fq∗. Notice that if we can write α as a sum of n + 1 many k powers, then any element in the coset of α, i.e. in αR , can be written as a sum of n + 1 many kth powers. This implies if α (n + 1)R , k∗ ∈ k∗ then αR (n + 1)R . This means the number of the strict inclusions in R 2R mR is at k∗ ⊆ k∗ k ⊆ k ⊆···⊆ k WARING’S PROBLEM IN FINITE RINGS 15 most one less than the number of cosets of R inside F . Moreover, since F is cyclic of order q 1 and R k∗ q∗ q∗ − k∗ k q 1 is a subgroup of Fq∗, Rk∗ has to be a . Let Fq∗ = γ , then Rk∗ = γ and Rk∗ = gcd(k,q− 1) . So, h i h i | | − the number of cosets is gcd(k,q 1). Also, it is clear that F = R if and only if gcd(k,q 1) = 1.  − q∗ k∗ −

This section proceeds like this. First, we prove that when a finite field is sufficiently large, every element of the field can be written as a sum of two kth powers, three kth powers etc. There exists some similar old results in the literature implying our result in [14], [19], [21]; but we will provide as elementary and self-contained proof as possible by following the graph theoretical approach. Our proof actually turns out to be very fruitful, since we also obtain an original result which is an analog of Sárközy’s theorem in a finite field setting. After Sárközy’s we provide the precise values for γ(k,q) for 4 6 k 6 37 for all finite fields (only except some of the γ(k,q) 6 3). Those results were obtained with the help of a supercomputer.

4 th Theorem 3.2. If q > k , then every element of Fq can be written as a sum of two k powers.

Proof. We will use the spectral gap theorem to prove the result. Since v = 0 is already a sum of two kth powers, let v F . We have d = gcd(k,q 1) many cosets of R inside F and v vR . The plan is to ∈ q∗ − k∗ q∗ ∈ k∗ take X = Rk∗, Y = vRk∗ and apply the spectral gap theorem. When we apply it, the theorem will guarantee the existence of a directed edge from X to Y under the assumption q > k4. This means there exists some uk X and vwk Y for some u, w, z F such that uk + zk = vwk i.e. v = ( u )k + ( z )k, and the result ∈ ∈ ∈ q∗ w w follows. The only thing left is to figure out when the spectral gap theorem applies.

First, define a Cayley digraph Cay(Fq,Rk∗) whose vertices are the elements of Fq, and there exists a directed edge from u to v if and only if v u R . Since R does not contain 0, Cay(F ,R ) is loopless. We − ∈ k∗ k∗ q k∗ know every character on Fq corresponds to an eigenvector of the adjacency matrix, A of Cay(Fq,Rk∗) by Theorem 2.1. For instance if F = γ we have F = 0, γ, γ2, , γq 1 = 1 , and any character χ gives q∗ h i q { · · · − } us a unique eigenvector such that the first entry of the corresponding eigenvector is χ(0), the second entry of the eigenvector is χ(γ), the third entry is χ(γ2) etc. We can even find the eigenvalues corresponding to those eigenvectors using the same theorem. Recall that we proved every character on Fq should be in the form of 2πix χ φ for some α F where φ (x) = tr(αx) and χ (x) = e p for every x F . Let λ denote the 1 ◦ α ∈ q α 1 ∈ q α eigenvalue of A corresponding to the eigenvector induced from χ φ . By Theorem 2.1 we have 1 ◦ α 2πi(tr(αx)) λ = χ φ (x)= e p . α 1 ◦ α x R∗ x R∗ X∈ k X∈ k Notice if αRk∗ = βRk∗ we have 2πi(tr(αx)) 2πi(tr(βx)) e p = e p = λβ. x R∗ x R∗ X∈ k X∈ k This calculation shows us if α and β are in the same coset of Rk∗ inside Fq∗, the eigenvectors induced by χ1 φα and χ1 φβ correspond to the same eigenvalue. We also have the trivial character which induces 1 ◦ ◦ F∗ | q | as an eigenvector with eigenvalue Rk∗ , this means we have at most R∗ + 1 many distinct eigenvalues. Thus | | | k| 16 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT the spectrum of A contains the largest eigenvalue R with multiplicity 1 and other eigenvalues, λ ’s each | k∗| αi with multiplicity R , since α R = R . Therefore, we have | k∗| | i k∗| | k∗| d 2 2 2 λ = R∗ + λ R∗ . (3) k αk | k| k αi k | k| α Fq i=1 X∈ X At the same time we know 2 λ = tr(A∗A) (4) k αk α Fq X∈ from basic linear algebra. Notice that we also have

T T + tr(A∗A) = tr(A A) = tr(AA )= d (α) (5) α Fq X∈ T as A is real A∗ = A , and the last equality holds since Cay(Fq,Rk∗) is a simple digraph. By Equations (3), (4) and (5) we have d 2 2 R∗ + λ R∗ = q R∗ . | k| k αi k | k| | k| Xi=1 Hence d 2 λ = q R∗ k αi k − | k| Xi=1 which implies 6 max λαi q Rk∗ . 16i6d k k − | | q It follows that

q q q Rk∗ n = max λαi 6 − | |. ∗ R 16i6d k k R | k∗| p| k∗|  Now, let x = d = gcd(k,q 1) and y = q in Lemma A.1 in Appendix A. Notice that since y is the order − of the finite field, it is already in Z 1 . Besides, there is no harm assuming x = 1, since we already + \ { } 6 handled x = 1 case before in Proposition 3.1. By the lemma, we have (y 1)4 x4y3 + (y 1)y2x3 > 0, (y 1)4 (y 1)y2 − − − by dividing it with x4 we get − + − >y3. This means R 4 + R q2 >q3 which is to say x4 x | k∗| | k∗| 4 3 2 4 2 4 R∗ >q R∗ q i.e. R∗ >q (q R∗ ). So q > k implies R∗ >n and the result follows.  | k| − | k| | k| − | k| | k| ∗

3 th Theorem 3.3. If q > k , then every element of Fq can be written as a sum of three k powers.

Proof. We will use Corollary 2.3 and the discussion provided right after that. We define a new digraph using the Cayley digraph in the previous theorem’s proof. The vertices of the new digraph are the elements of

Fq, and there exists a directed edge from u to v if and only if there is a 2-walk from u to v in the original graph. Since v = 0 is already a sum of three kth powers, let v F . Let X = R , Y = vR and apply ∈ q∗ k∗ k∗ Corollary 2.3. It will guarantee the existence of a directed 2-walk from X to Y under the assumption q > k3 and the result will follow. WARING’S PROBLEM IN FINITE RINGS 17

Let x = d = gcd(k,q 1) and y = q in Lemma A.2 in Appendix A. By the lemma, we have (y 1)3 − 3 − − 2 3 (y 1) xy y+1 3 x y(xy y + 1) > 0, by dividing it with x we get −x3 >y −x . This means Rk∗ >q(q Rk∗ ) − ∗ | | − | | q(q Rk ) −|∗ |    i.e. Rk∗ > R 2 and the result follows. | | | k| − m 1 2m m m+1 m 1 Remark. If one can prove that y > x implies (y 1) yx 2 (xy y + 1) 2 > 0 for every − − − − m,x,y Z 1 , then using the same exact approach in the previous theorems’ proof, we can state this ∈ + \ { } 2m q > k m−1 = F mR ⇒ q ⊆ k − m m+1 m 1 for every integer m > 1. The required inequality (y 1) yx 2 (xy y + 1) 2 > 0 whenever − − − m 1 2m y − > x is proven for m = 2 and m = 3 in Appendix A for the proofs of the previous theorems; but we did not prove it here for all m Z 1 . This assertion is actually already familiar to us, since it is ∈ + \ { } asymptotically equivalent to Winterhof’s result. Recall Winterhof proved that

2m q > (k 1) m−1 = F mR . − ⇒ q ⊆ k He used Jacobi sums to deduce his bound, see [21].

Let the notation stay the same with the last two theorems. Then, we have another nice result:

Theorem 3.4 (Analog of Sárközy’s Theorem in Finite Fields). Let k be a positive integer. If E is a subset qk of F with size E > then it contains at least two distinct elements whose difference is a kth power. q | | √q 1 − Thus, in particular if k is fixed and E =Ω √q as q , then E contains at least two distinct elements | | →∞ th whose difference is a k power. 

Proof. As in the last two theorems, again we want to apply the spectral gap theorem. Our calculations in the q q R proof of Theorem 3.2 showed that n 6 − | k∗|. We have ∗ R p| k∗| q 1 q q R q q −d q − | k∗| = − = √d qd q + 1. R q q 1 q 1 − p k∗ −d | | − p Since d = gcd(k,q 1) 6 q 1 i.e. d + 1 6 q we have − − q q qd qk √d qd q + 1 6 √d d(q 1) = 6 . q 1 − q 1 − √q 1 √q 1 − p − p − − If qk , by Theorem 2.2 there is an edge from a vertex in to a distinct vertex in and the result √q 1 < E E E − | | follows. 

For example consider F412 = F1681. If we take an arbitrary subset of F1681 with size greater than or equal to 124 (165, 206 resp.), then we know there exists at least two distinct elements in this set whose difference is a 3rd (4th, 5th resp.) power. 18 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

We claimed before that we will state the precise values of γ(k,q) for every 4 6 k 6 37 only except some of γ(k,q) 6 3. Hence let 4 6 k 6 37 be given. First note that by the theory, we know that for every k, there exists only a finite number of finite fields which are not coverable with the kth powers, see e.g. [17]. Table 1 lists all of the uncoverable finite fields for 4 6 k 6 37. If we are given with a specific q as well, then we can determine if γ(k,q) = 1 using the criteria in Proposition 3.1. Theorem 3.6 and 3.7 (upcoming) handle all of the cases where γ(k,q) = 4, 5, 6 and some of the cases where γ(k,q) = 3. Theorem 3.8 provides all of the fields such that γ(k,q) > 7. Therefore, if we pick a 4 6 k 6 37 and a coverable field Fq by taking into consideration Table 1, and if it doesn’t fall into one of the categories in Theorem 3.2, 3.6, 3.7 or 3.8, then we can conclude γ(k,q) 6 3.

The results in Proposition 3.5 below are discovered via exhaustive search using Sage, in principle one can calculate these with Theorem G in [3] doing finitely many calculations.

Proposition 3.5. Table 1 lists all of the uncoverable fields for any 4 6 k 6 37.

TABLE 1. Uncoverable Fields

k uncoverable fields for k k uncoverable fields for k 4 9 20 9, 16, 81, 361 5 16 21 4, 8, 64 6 4, 25 24 4, 9, 25, 49, 121, 529 7 8 25 16 8 9, 49 26 27, 625 9 4, 64 27 4, 64 10 16, 81 28 8, 9, 169, 729 12 4, 9, 25, 121 30 4, 16, 25, 81, 841 13 27 31 32, 125 14 8, 169 32 9, 49, 961 15 4, 16 33 4, 1024 16 9, 49 34 256 17 256 35 8, 16 18 4, 25, 64, 289 36 4, 9, 25, 64, 121, 289

Notice that k = 11, 19, 22, 23, 29, 37 cases are not on the Table 1. It is because there are not any uncoverable fields corresponding to those k values.

Theorem 3.6. For every integer k satisfying 4 6 k 6 19, Table 2 lists all of the fields with size q such that γ(k,q) = 3, 4, 5 or 6.

Table 2: Fields

k γ(k, q) = 3 γ(k, q) = 4 γ(k, q) = 5 γ(k, q) = 6 4 13, 17, 25, 29 5 5 31, 41, 61 11 6 37, 43, 49, 61, 67, 73, 79, 109, 139, 223 19, 31 7, 13 WARING’S PROBLEM IN FINITE RINGS 19

Table 2: (continued)

k γ(k, q) = 3 γ(k, q) = 4 γ(k, q) = 5 γ(k, q) = 6 7 64, 71, 113, 127 29, 43 8 13, 29, 73, 81, 89, 97, 113, 121, 137, 233, 257, 289, 337, 5, 25, 41 761 9 7, 73, 109, 127, 163, 181, 199, 271, 307, 343 37 10 71, 101, 121, 131, 151, 181, 191, 211, 241, 251, 271, 281, 41, 61 31 311, 331, 401, 421, 431, 461, 491, 641, 911 11 199, 331, 353, 419, 463, 617 89 67 12 17, 29, 43, 67, 79, 139, 157, 169, 181, 193, 223, 241, 277, 5, 19, 31, 61, 73 7, 37, 49 289, 313, 337, 349, 361, 373, 397, 409, 421, 433, 457, 541, 97, 109, 229 577, 625, 661, 673, 733, 841, 877, 1069, 1453, 1669, 1741 13 131, 157, 313, 443, 521, 547, 599, 677, 859, 911, 937, 79 53 1171 14 64, 197, 211, 239, 281, 337, 379, 421, 449, 463, 491, 547, 71, 113, 127 43 617, 631, 659, 673, 701, 743, 757, 827, 911, 953, 967, 1009, 1499, 2927 15 7, 41, 151, 181, 211, 241, 256, 271, 331, 361, 421, 541, 121 11, 61 571, 601, 631, 661, 691, 751, 811, 961, 991, 1021, 1051, 1171, 1201, 3181 16 13, 29, 73, 89, 121, 137, 233, 241, 289, 337, 353, 401, 433, 5, 25, 41, 81, 97 449, 529, 577, 593, 625, 641, 673, 761, 769, 881, 929, 977, 113, 193, 257 1009, 1201, 1249, 1297, 1409, 1489, 1697, 2017, 2401, 2593 17 239, 307, 409, 443, 613, 647, 919, 953, 1021, 1123, 1259, 137 103 1327, 1361, 1531, 1667 18 43, 49, 61, 67, 79, 139, 223, 307, 343, 361, 379, 433, 487, 31, 163, 181, 199 7, 13, 109, 523, 541, 577, 613, 631, 739, 757, 811, 829, 883, 919, 271, 397 127 937, 991, 1009, 1063, 1117, 1153, 1171, 1279, 1369, 1423, 1459, 1693, 1747, 2089, 2197, 2251, 2269, 2287, 2503, 2719, 3259, 4519 19 343, 457, 571, 647, 761, 1103, 1217, 1483, 1559, 1597, 191, 229, 419 1787, 2053, 2129, 2357, 2927

Theorem 3.7. Let 20 6 k 6 37 be any given integer.

Table 3 lists all of the fields with size q such that γ(k,q) = 4, 5 or 6. • Table 3 lists all of the fields with size q such that γ(k,q) = 3 and q 6 (k 1)3. • − (We do not list all of the finite fields with γ(k,q) = 3 here, since when we ran the computer, we used a bound from [21], which is if q > (k 1)3 then F 3R .) − q ⊆ k 20 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

Table 3: Fields

k γ(k, q) = 3 γ(k, q) = 4 γ(k, q) = 5 γ(k, q) = 6 20 13, 17, 25, 29, 71, 131, 151, 191, 211, 251, 271, 281, 311, 5, 181, 241, 521 31, 101 331, 401, 421, 431, 461, 491, 541, 601, 641, 661, 701, 761, 821, 841, 881, 911, 941, 961, 1021, 1061, 1181, 1201, 1301, 1321, 1361, 1381, 1481, 1601, 1621, 1681, 1721, 1741, 1861, 1901, 2221, 2281, 2381, 2621, 2861 21 7, 71, 113, 379, 421, 463, 547, 631, 673, 757, 841, 883, 29, 169, 211, 127 967, 1009, 1051, 1093, 1303, 1429, 1471, 1597, 1723, 337 1849, 2143, 2437, 2521, 2647, 2689, 2857, 3319, 5167 22 243, 353, 397, 529, 617, 661, 683, 727, 859, 881, 947, 991, 199, 331, 419, 1013, 1123, 1277, 1321, 1409, 1453, 1607, 1783, 1871, 463 2003, 2069, 2113, 2179, 2267, 2333, 2377, 2399, 2531, 2663, 2729, 2927, 3037, 3719, 3851, 3917 23 461, 599, 691, 829, 967, 1013, 1151, 1289, 1381, 1427, 277 1657, 1933, 1979, 2209, 2347, 2393, 2531, 3083, 3313, 4831, 4969, 5659, 8419 24 29, 43, 67, 79, 81, 89, 113, 137, 139, 157, 181, 223, 233, 5, 19, 31, 41, 193, 289 7, 37 257, 277, 337, 349, 373, 397, 421, 433, 541, 601, 625, 661, 61, 109, 169, 673, 733, 761, 769, 841, 877, 937, 961, 1009, 1033, 1069, 229, 241, 313, 1129, 1153, 1201, 1249, 1297, 1321, 1369, 1453, 1489, 361, 409, 457, 1609, 1657, 1669, 1681, 1741, 1753, 1777, 1801, 1849, 577 1873, 1993, 2017, 2089, 2113, 2137, 2161, 2281, 2377, 2401, 2473 25 31, 41, 61, 401, 601, 701, 751, 1051, 1151, 1201, 1301, 251 11 1451, 1601, 1801, 1901, 1951, 2251, 2351, 2551, 2801, 2851, 3001, 3251, 3301, 3851, 5051, 5501 26 521, 677, 729, 859, 911, 937, 1093, 1171, 1223, 1249, 313, 443, 547, 131 1301, 1327, 1483, 1613, 1847, 1873, 1951, 2003, 2029, 599 2081, 2237, 2341, 2393, 2549, 2809, 2861, 2887, 2939, 3121, 3329, 3407, 3433, 3511, 3719, 3797 27 7, 73, 127, 181, 199, 307, 343, 379, 757, 811, 919, 1297, 271, 433, 487, 37 1459, 1567, 1621, 1783, 1999, 2053, 2161, 2269, 2377, 541 2539, 2593, 2647, 2917, 2971, 3079, 3187, 3457, 3727, 3889, 3943, 4051, 4159, 4483, 4591, 4861, 4969, 5023, 6859, 7993, 9397, 9829 28 13, 17, 25, 64, 211, 239, 379, 463, 491, 547, 631, 659, 673, 5, 71, 127, 337, 197, 281 43 701, 743, 757, 827, 841, 911, 953, 967, 1009, 1093, 1289, 421, 449, 617 1373, 1429, 1499, 1597, 1681, 1709, 1849, 1877, 1933, 2017, 2129, 2213, 2269, 2297, 2381, 2437, 2521, 2549, 2633, 2689, 2801, 2857, 2927, 2969, 3109, 3137, 3221, 3361, 3389, 3529, 3557, 3613, 3697, 4201, 4229, 4397, 4481, 4621, 4649, 4733, 4789, 4817, 4957 29 523, 929, 1103, 1277, 1451, 1567, 1741, 1973, 2089, 2437, 349 233 3191, 3307, 3481, 3539, 4003, 4177, 4409, 4583, 4931, 5569, 5801, 6091, 8237 WARING’S PROBLEM IN FINITE RINGS 21

Table 3: (continued) k γ(k, q) = 3 γ(k, q) = 4 γ(k, q) = 5 γ(k, q) = 6 30 37, 43, 49, 67, 71, 73, 79, 101, 109, 131, 139, 191, 223, 19, 41, 271, 211, 241 7, 13, 151 251, 256, 281, 311, 401, 431, 461, 491, 601, 631, 641, 331, 361, 421, 691, 751, 811, 911, 961, 991, 1021, 1051, 1171, 1201, 541, 571, 661 1231, 1291, 1321, 1381, 1471, 1531, 1621, 1681, 1741, 1801, 1831, 1861, 1951, 2011, 2131, 2161, 2221, 2251, 2281, 2311, 2341, 2371, 2401, 2521, 2551, 2671, 2731, 2791, 2851, 2971, 3001, 3061, 3121, 3181, 3271, 3301, 3331, 3361, 3391, 3511, 3541, 3571, 3631, 3691, 3721, 3931, 4021, 4051, 4111, 4201, 4261, 4441, 4561, 4831, 4861 31 683, 1024, 1117, 1303, 1427, 1489, 1613, 1861, 2357, 311 373 2543, 2729, 2791, 3163, 3659, 3907, 4093, 4217, 4651, 5023, 5147, 5209, 5519, 5581, 5953, 7193 32 13, 29, 73, 89, 121, 137, 233, 241, 337, 401, 433, 529, 5, 25, 41, 81, 257 593, 625, 641, 673, 761, 881, 929, 977, 1009, 1153, 1201, 113, 289, 353, 1217, 1249, 1297, 1409, 1489, 1601, 1697, 1889, 2017, 449, 577, 769 2081, 2113, 2209, 2273, 2401, 2593, 2657, 2689, 2753, 3041, 3137, 3169, 3329, 3361, 3457, 3617, 4001, 4129, 4289, 4513, 4673, 4801, 4993 33 7, 353, 419, 617, 859, 991, 1123, 1321, 1453, 1783, 1849, 89, 397, 463, 331 2113, 2179, 2311, 2377, 2707, 2971, 3037, 3169, 3301, 529, 661, 727 3433, 3499, 3631, 3697, 4027, 4093, 4159, 4357, 4423, 4489, 4621, 4951, 5281, 5347, 5413, 5479, 5743, 6007, 6073, 6271, 6337, 6469, 6997, 7723, 7789, 7921, 8317, 8647, 9439 34 919, 1021, 1123, 1259, 1327, 1361, 1429, 1531, 1667, 443, 613, 647, 307, 409 239 1871, 1973, 2143, 2347, 2381, 2551, 2687, 2789, 2857, 953 3061, 3163, 3299, 3469, 3571, 3673, 3877, 3911, 4013, 4217, 4421, 4523, 4591, 4931, 4999, 5101, 5237, 5407, 5441, 5849 35 31, 41, 61, 64, 113, 127, 701, 841, 911, 1051, 1331, 1471, 29, 43, 421, 11, 281 1681, 2311, 2381, 2521, 2591, 2731, 2801, 3011, 3221, 491, 631 3361, 3571, 3851, 4096, 4201, 4271, 4481, 4621, 4691, 4831, 5041, 5531, 5741, 5881, 6301, 6581, 6791, 7001, 7351, 7561, 8191, 8681, 8821 36 17, 29, 43, 67, 79, 139, 157, 169, 193, 223, 241, 277, 307, 5, 31, 61, 97, 199, 361, 397 7, 49, 127 313, 337, 343, 349, 373, 379, 409, 421, 457, 487, 523, 163, 229, 271, 625, 631, 661, 673, 733, 739, 811, 829, 841, 877, 883, 433, 541, 577, 919, 937, 991, 1009, 1063, 1069, 1153, 1171, 1279, 1297, 613, 757, 1117 1369, 1423, 1453, 1459, 1549, 1621, 1657, 1669, 1693, 1741, 1747, 1801, 1873, 2017, 2053, 2089, 2161, 2197, 2251, 2269, 2287, 2341, 2377, 2503, 2521, 2557, 2593, 2719, 2809, 2917, 2953, 3061, 3169, 3259, 3313, 3457, 3529, 3637, 3673, 3709, 3853, 3889, 4177, 4357, 4519, 4789, 4861, 4933, 4969 22 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

Table 3: (continued)

k γ(k, q) = 3 γ(k, q) = 4 γ(k, q) = 5 γ(k, q) = 6 37 1259, 1481, 1777, 1999, 2221, 2591, 2887, 3109, 3257, 593 3331, 3701, 3923, 4219, 4441, 4663, 5107, 5477, 6143, 6217, 6661, 7253, 7549, 7919, 7993, 8363, 8807, 9103, 9473

Theorem 3.8. For 4 6 k 6 37, the following is the list of all of the fields with size q such that γ(k,q) > 7 :

γ(8, 17) = 8 • γ(9, 19) = 9 • γ(10, 11) = 10 • γ(11, 23) = 11 • γ(12, 13) = 12 • γ(14, 29) = 14 • γ(15, 31) = 15 • γ(16, 17) = 16 • γ(18, 73) = 7, γ(18, 19) = 18, γ(18, 37) = 18 • γ(20, 121) = 7, γ(20, 61) = 8, γ(20, 11) = 10, γ(20, 41) = 20 • γ(21, 43) = 21 • γ(22, 89) = 7, γ(22, 67) = 8, γ(22, 23) = 22 • γ(23, 139) = 7, γ(23, 47) = 23 • γ(24, 17) = 8, γ(24, 73) = 8, γ(24, 97) = 8, γ(24, 13) = 12 • γ(25, 151) = 7, γ(25, 101) = 9 • γ(26, 157) = 8, γ(26, 79) = 9, γ(26, 53) = 26 • γ(27, 163) = 8, γ(27, 19) = 9, γ(27, 109) = 9 • γ(28, 113) = 7, γ(28, 29) = 28 • γ(29, 59) = 29 • γ(30, 181) = 8, γ(30, 11) = 10, γ(30, 121) = 10, γ(30, 31) = 30, γ(30, 61) = 30 • γ(32, 193) = 8, γ(32, 97) = 10, γ(32, 17) = 16 • γ(33, 199) = 9, γ(33, 23) = 11, γ(33, 67) = 33 • γ(34, 103) = 10, γ(34, 137) = 10 • WARING’S PROBLEM IN FINITE RINGS 23

γ(35, 211) = 9, γ(35, 71) = 35 • γ(36, 181) = 7, γ(36, 109) = 11, γ(36, 13) = 12, γ(36, 19) = 18, γ(36, 37) = 36, • γ(36, 73) = 36

γ(37, 149) = 9, γ(37, 223) = 9. •

Let k be any positive integer. Then, we define γ(k) := max γ(k,q) where we take the maximum value of γ(k,q) over the set q γ(k,q) exists . Hence, if a finite field is coverable with the kth powers, every { | } element of the field can be written as a sum of γ(k) many kth powers. By Proposition 3.1 we have γ(k) 6 k. The following proposition gives us the exact values of γ(k) for k 6 37:

Proposition 3.9. Let k 6 37. If γ(k) is not provided in Table 4, that means γ(k)= k.

TABLE 4. γ(k) values

γ(7) = 4 γ(13) = 6 γ(17) = 6 γ(19) = 4 γ(24) = 12 γ(25) = 9 γ(27) = 9 γ(31) = 5 γ(32) = 16 γ(34) = 10 γ(37) = 9

4. WARING’S PROBLEMIN MATRIX RINGS

Let Fq be the finite field with q elements and let p denote the characteristic of Fq. Also let k be a positive integer and Mat (F ) be the ring of n n matrices over F . In this section we want to find an integer m n q × q th with the property that every element of Matn(Fq) can be written as a sum of at most m many k powers in Mat (F ). First, we need to develop our machinery. Take any A Mat (F ). Let P (x) F [x] be n q ∈ n q ∈ q any polynomial over F , then P (x) = a + a x + + a xl for some a , a a F . We can evaluate q 0 1 · · · l 0 1 · · · l ∈ q any polynomial P (x) at A; for example here P (A) = a I + a A + + a Al where I denotes the n n 0 1 · · · l × identity matrix over F . Then, we define F [A]= P (A) P (x) F [x] . We want to find an m such that q q { | ∈ q } A = Bk + + Bk for some B , ,B Mat (F ). Instead of working with Mat (F ) we will work 1 · · · m 1 · · · m ∈ n q n q th with Fq[A], meaning we will try to write A as a sum of k powers using the elements of Fq[A]. This is more convenient since we have more tools to use in Fq[A], but we are only able to provide upper bounds for m with this approach, rather than exact values. Define

ϕ : F [x] F [A] q −→ q a + a x + + a xl a I + a A + + a Al. 0 1 · · · l 7−→ 0 1 · · · l Notice that

ker ϕ = P (x) Fq[x] s.t. P (A) = 0n n = mA(x) { ∈ × } h i where mA(x) denotes the minimum polynomial of A, and Im ϕ = Fq[A]. By first theorem we have F q[x] m (x) = Fq[A]. h A i ∼ . 24 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

Let’s assume m (x)= p (x)i1 p (x)ij for some p ,p , ,p distinct irreducible polynomials over F . A 1 · · · j 1 2 · · · j q Using Chinese remainder theorem we have

Fq[x] pi1 pij = Fq[x] pi1 Fq[x] pij . h 1 · · · j i ∼ h 1 i ×···× h j i . . . We need one more tool before we state the main result of the section.

Proposition 4.1. Let f(x) be an irreducible polynomial over Fq with degree n. Let i and k be positive integers such that p ∤ k where p denotes the characteristic of Fq.

If every element of Fq[x] f(x) is a kth power, then the elements of Fq[x] f i(x) can be written • h i h i as a sum of two kth powers.. .

Assume every element of Fq[x] f(x) can be written as a sum of m many kth powers for some • h i th . m> 1. If 1 is a k power in Fq, then the elements of Fq[x] f i(x) can be written as a sum of m − h i th th . many k powers. If 1 is not a k power in Fq, then the elements of Fq[x] f i(x) can be written − h i as a sum of m + 1 many kth powers. .

Proof. We denote f(x) with f in this proof. Take any element from Fq[x] f 2 . Using division algorithm h i we can write an equivalence class representative of this element as r +.r f where r , r F [x] and 1 2 1 2 ∈ q deg(r ), deg(r ) < n. It follows from the assumption that there exist some B ,B , ,B F [x] such 1 2 1 2 · · · m ∈ q that r Bk + Bk + + Bk (mod f). 1 ≡ 1 2 · · · m

If some of Bi’s are zeros modulo f, we can erase them from the list of Bi’s, and use the ones which are not zero modulo f for the next step. We can erase them, because if B 0 (mod f), Bk 0 (mod f) so the i ≡ i ≡ th summation of the k powers of the reduced list will still be r1. If all of the Bi’s are zeros modulo f, then we need to be a little bit careful. If k is odd, or if k is even and 1 is a kth power in F , simply take B = 1 − q i1 and B = 1 and proceed to the next step. If k is even and 1 is not a kth power in F , then we can write i2 − − q 0 Bk + ( 1) (mod f) for B = 1, and we know 1 can be written as a sum of m many kth powers by ≡ i1 − i1 − assumption. So we can assume B , ,B are all nonzero in modf and their summation is equivalent to i1 · · · ij r1 in modf.

k k k Define the polynomial Q(t) = t + B + + B r1 r2f over Fq[x], use Hensel’s lemma for the i1 · · · ij−1 − − k k polynomial rings (Proposition 2.16) for Q(t) to conclude there exists a g2 Fq[x] such that g + B + + ∈ 2 i1 · · · Bk = r + r f (mod f 2) and the result for i = 2 follows. Here, we are allowed to use Hensel’s lemma, ij−1 1 2 1 k 1 since Bij is a root of Q(t) in modf and ν(Q′(Bij )) = 0 < 2 . ν(Q′(Bij )) = 0 since Q′(t) = kt − and k 1 Q′(B )= kB − 0 (mod f). This follows from f ∤ k, p ∤ k, f ∤ B and f being irreducible. ij ij 6≡ ij

Once we lift a solution from Fq[x] f(x) to Fq[x] f 2(x) , it is easy. Third condition in Proposition 2.16 h i h i allows us to apply Hensel’s lemma. again and again,. and get a solution in each Fq[x] f i(x) for any h i i Z . .  ∈ + WARING’S PROBLEM IN FINITE RINGS 25

Using this proposition, the discussion right before the proposition and finite field tables in Section 3 we literally have dozens of nice results for the matrix rings, but it would be a waste of paper to write out all of them. That is why we will state the results only for some small k values to get across our point to the th reader in the following theorem. Note that we denote with Rk,n the set of all k powers in Matn(Fq) i.e. R = M k M Mat (F ) . Also, Mat (F ) mR means every element of Mat (F ) can be k,n | ∈ n q n q ⊆ k,n n q th written as a sum of m many k powers. Theorem 4.2. In the following table in a fixed row, let k be a positive integer as in the first column, and let Fq be the finite field with q elements where q is different than the values in the second column. Then Mat (F ) mR for any n. In fact, in this case any A Mat (F ) has A = Bk + Bk + + Bk n q ⊆ k,n ∈ n q 1 2 · · · m with B , B , ,B in the subring generated by A. (For example the first row should be read as "If 1 2 · · · m F = 2, 4, or any power of 3, then Mat (F ) 3R for any n. In fact, in this case any A Mat (F ) | q| 6 n q ⊆ 3,n ∈ n q 3 3 3 has A = B1 + B2 + B3 with B1, B2, B3 in the subring generated by A.")

k = q =6 m = 3 2, 4, 3s 3 2, 4, 7, 3s 2 4 3, 9, 2s 5 3, 5, 9, 2s 4 3, 5, 9, 13, 17, 25, 29, 2s 3 5 2, 4, 16, 5s 5 2, 4, 11, 16, 5s 3 2, 4, 11, 16, 31, 41, 61, 5s 2 6 2, 4, 5, 25, 2s, 3s 7 2, 4, 5, 7, 13, 25, 2s, 3s 5 2, 4, 5, 7, 13, 19, 25, 31, 2s, 3s 4 2, 4, 5, 7, 13, 19, 25, 31, 37, 43, 49, 61, 67, 73, 79, 109, 139, 223, 2s, 3s 3 7 2, 8, 7s 4 2, 8, 29, 43, 7s 3 2, 4, 8, 29, 43, 64, 71, 113, 127, 7s 2 8 3, 7, 9, 49, 2s 9 3, 7, 9, 17, 49, 2s 5 3, 5, 7, 9, 17, 25, 41, 49, 2s 4 3, 5, 7, 9, 11, 13, 17, 25, 29, 41, 49, 73, 81, 89, 97, 113, 121, 137, 233, 257, 3 289, 337, 761, 2s 9 2, 4, 8, 64, 3s 9 2, 4, 8, 19, 64, 3s 5 2, 4, 8, 19, 37, 64, 3s 3 2, 4, 7, 8, 19, 37, 64, 73, 109, 127, 163, 181, 199, 271, 307, 343, 3s 2 10 2, 3, 4, 9, 16, 81, 2s, 5s 11 2, 3, 4, 9, 11, 16, 81, 2s, 5s 6 2, 3, 4, 9, 11, 16, 31, 81, 2s, 5s 5 2, 3, 4, 9, 11, 16, 31, 41, 61, 81, 2s, 5s 4 2, 3, 4, 9, 11, 16, 31, 41, 61, 71, 81, 101, 121, 131, 151, 181, 191, 211, 241, 3 251, 271, 281, 311, 331, 401, 421, 431, 461, 491, 641, 911, 2s, 5s 26 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

11 23, 11s 5 23, 67, 11s 4 23, 67, 89, 11s 3 23, 67, 89, 199, 331, 353, 419, 463, 617, 11s 2

Proof. To explain the process, we will only do for the first row. Let k = 3. By Section 3 we know that every element of a finite field can be written as a sum of three cubes as long as q = 4. Let A Mat (F ). By 6 ∈ n q i ij the discussion prior to Proposition 4.1, we know that Fq[A] = Fq[x] p 1 Fq[x] p for some ∼ h 1 i ×···× h j i p ,p , ,p distinct irreducible polynomials over F . We first exclude. all of the subfields. of F from our 1 2 · · · j q 4 list so that none of the Fq[x] p will be isomorphic to F4. So, Fq[x] p will be different than F4, as h ri h ri . . long as Fq = F2, F4. This guarantees that every element of Fq[x] p can be written as a sum of three 6 h ri th . cubes. Then, we want to lift the k powers from Fq[x] p to Fq[x] pir . We will use Proposition 4.1. h ri h r i To use this proposition we need the characteristic of the. finite field not. to divide k. That is why we need to exclude p = 3 cases. Once we guarantee that three cubes is enough for Fq[x] pir for every 1 6 r 6 j, h r i then Chinese remainder theorem guarantees that three cubes is enough for Fq[x]. pi1 Fq[x] pij . h 1 i ×···× h j i One last thing to note is that when we determine m, if 1 is a kth power (in particular. if k is odd). we can − use the same m from the downstairs Fq[x] p ; but if 1 is not a kth power (for example in some cases h ri − when k is even) we need to increasem by .1 when we do the lifting using Proposition 4.1. That is why in the table when k is even, we increased m by 1 just to be safe without paying any attention to 1 being a kth − power or not. 

Note if we have a finite semisimple ring, by Artin-Wedderburn theory it is a direct product of finitely many matrix rings over finite fields and we can use matrix rings results to get sharp bounds for the Waring problem on finite semisimple rings. For example if we know R is semisimple with R = 5411213, then we know | | every element of R can be written as a sum of two cubes; since there is not any possibility that R has a over F2, F4 or a finite field of characteristic 3 in its decomposition.

Next we give an example which shows that sometimes the matrix rings are better than the finite fields in terms of coverability, since we have more flexibility.

Example. Take F4 = F2[x] x2 + x + 1 so that we can denote the elements with 0, 1,x,x + 1 . We ∼ h i { } . 1 1 x + 1 1 have x2 = x + 1, (x + 1)2 = x2 +1= x and x(x + 1) = 1. Let A = and B = . 1 x + 1 1 1     x 1 0 1 x 0 Then, A3 = and B3 = . So, A3 + B3 = . This is very interesting although we 1 0 1 x 0 x       x 0 cannot write x as a sum of cubes in F (since x3 = (x + 1)3 = 1 in F ), we can write as a sum of 4 4 0 x   two cubes in Mat2(F4). Furthermore, one can also show that every element of Mat2(F4) can be written as a sum of two cubes. WARING’S PROBLEM IN FINITE RINGS 27

5. WARING’S PROBLEMIN FINITE RINGS

Let R be a finite ring with identity which is not necessarily commutative, and let k be any positive integer. In this section we want to find an integer n with the property that every element of R can be written as a sum of at most n many kth powers in R. Before we dive into results and proofs, first we need to setup our machinery for finite rings with identity. First notice R /J is semisimple i.e. J R /J = 0. Moreover, R R R since R is finite, /J is finite so /J is both left and right Artinian. Theorem 2.6 implies that /J ∼= R Matn1 (D1) Matnr (Dr) for some D1, Dr division rings. Since /J is finite, each Di has to ×···× · · · have finitely many elements. By Wedderburn’s little theorem Di’s are finite fields. Therefore, we have R /J = Matn1 (Fq1 ) Matnr (Fqr ) for some finite fields Fq1 , , Fqr . Furthermore ∼ ×···× · · · J J 2 J 3 J l = (0) ⊇ ⊇ ⊇···⊇ for some l by Corollary 2.9 and

R l ։ R l 1 ։ R 2 ։ R R = /J /J /J /J = Matn1 (Fq1 ) Matnr (Fqr ). − · · · ∼ ×···× Theorem 5.1. Let R be a finite ring with identity which is not necessarily commutative and J denote the Jacobson radical of R. Let k be a positive integer such that gcd( R , k) = 1 where R denotes the order of | | | | the ring.

th th If every element of R /J is a k power, then unit elements of R are k powers and nonunit elements • of R can be written as a sum of two kth powers.

Let m> 1. Assume every element of R /J can be written as a sum of m many kth powers in R /J . • If α R is a unit (or equivalently if α mod J in R /J is a unit), then α can be written as a sum of ∈ m many kth powers in R. If k is odd (resp. even), then every element of R can be written as a sum of m (resp. m + 1) many kth powers in R.

R Proof. Case 1: Let R be commutative. Then we have /J = Fq1 Fqr for some finite fields, Fqi ’s. ∼ ×···× Let α R. Since we assumed every element of R /J can be written as a sum of m many kth powers in ∈ R k k R /J , we know α B + + Bm mod J for some B1, ,Bm R. We can send α from R to /J ≡ 1 · · · · · · ∈ with the reduction mod J, and denote it with α¯ so that α¯ = (α , α , , α ) for some α F where 1 2 · · · r i ∈ qi 1 6 i 6 r. Similarly, we will use B¯ to denote B mod J for every B R. j j j ∈

Case 1.a: If α¯ is a unit in R /J that means each entry of α¯ (i.e. αi for every 1 6 i 6 r) is nonzero. Then we can arrange B¯j’s (by shuffling the entries of B¯j’s) such that B¯1 does not have any zero entry, so that B¯1 ¯ is a unit in Fq1 Fqr . Since we have gcd( R , k) = 1 by assumption, k = (k1, k2, , kr) that is the ×···× | | k 1 · · · image of k under reduction mod J is a unit in F F . This implies k¯B¯ − is a unit, and the result q1 ×···× qr 1 follows from Corollary 2.14.

Case 1.b: If α¯ is not a unit in R /J , that means at least one of the entries of α¯ is zero. Assume only one of them equals 0, say α . We have α Bk + + Bk mod J by assumption. We can shuffle the entries 1 ≡ 1 · · · m 28 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

th of B¯j’s such that the i entry of B¯1 is nonzero for every 2 6 i 6 r. If the first entry of one of the B¯j’s is nonzero, then we can shuffle the nonzero entry to B¯1 and B¯1 becomes again a unit, we are back in case

1.a. If the first entries of all of the B¯j’s are zero for 1 6 j 6 m, then we can do the following trick. We ¯ ¯ ¯k can replace 0 in the first entry of B1 with 1, so that B1 becomes a unit. But then the first entry of B1 also becomes 1 and we need to destroy the effect of 1 in the summation (B¯k + B¯k + + B¯k =α ¯). Therefore, 1 2 · · · m we have to change the first entries of B¯j’s for 2 6 j 6 m such that the first entry of the summation B¯k + + B¯k will be 1. If k is odd, then simply replace 0 in the first entry of B¯ with 1, leave 2 · · · m − 2 − the other B¯ ’s as they are and the problem is solved since ( 1)k = 1. If k is even, by assumption j − − th R ( 1, 0, , 0) can be written as a sum of m many k powers in /J , this implies for some xj Fq1 we − · · · ∈ have ( 1, 0, , 0) = (x , 0, , 0)k +(x , 0, , 0)k + +(x , 0, , 0)k. Let B¯ = (x , 0, , 0), − · · · 1 · · · 2 · · · · · · m · · · new 1 · · · and replace the first entries of B¯j’s with xj’s for 2 6 j 6 r, so that the first entry of the summation ¯k ¯k ¯k ¯k ¯k ¯k ¯k Bnew + B2 + + Bm will be 1 and α¯ will be equal to B1 + Bnew + B2 + + Bm. We again have k 1 · · · − · · · k¯B¯ − is a unit, since gcd( R , k) = 1 and B¯ is a unit. Corollary 2.14 implies that α can be written as a 1 | | 1 sum of m + 1 many kth powers.

Case 2: Let R be noncommutative. To reduce this case to the commutative case, we use the following trick. Let α R. Consider the subring generated by α, i.e. Z[α]= a .1+ a α + + a αl a , a a Z . ∈ 0 1 · · · l | 0 1 · · · l ∈ It is a subring but it is a commutative finite ring with identity by itself. Moreover, since the order of Z[α] has to divide R , gcd( R , k) = 1 implies gcd( Z[α] , k) = 1 and we are again in case 1. Note that this approach | | | | | | enables us to write every element of Z[α] as a sum of kth powers from Z[α] instead of R, which means when R is not commutative the n value (i.e. the number of kth powers needed to write α as a sum) we find with this approach can be bigger than the actual n value. 

We actually proved a stronger result in the last proof. It follows that in the second part of Theorem 5.1 if 1 − is a kth power in R, then every element of R can be written as a sum of m many kth powers in R, otherwise we need to increase m by 1.

Theorem 5.2. Let R be a finite ring with identity which is not necessarily commutative. In the following table in a fixed row, let k be a positive integer as in the first column. If none of the q values stated in the second column divides the order of the ring, then every element of R can be written as a sum of n many kth powers in R. In fact, in this case any α R has α = Bk + Bk + + Bk with B , B , ,B in Z[α]. ∈ 1 2 · · · n 1 2 · · · n (For example the first row should be read as "If 3, 4 ∤ R , then any element of R can be written as a sum of | | three cubes in R. In fact, in this case any α R has α = B3 + B3 + B3 with B , B , B in Z[α].") ∈ 1 2 3 1 2 3

k = q ∤ n = 3 3, 4 3 3, 4, 7 2 4 2, 9 5 2, 5, 9 4 2, 5, 9, 13, 17, 29 3 WARING’S PROBLEM IN FINITE RINGS 29

5 5, 16 5 5, 11, 16 3 5, 11, 16, 31, 41, 61 2 6 2, 3, 25 7 2, 3, 7, 13, 25 5 2, 3, 7, 13, 19, 25, 31 4 2, 3, 7, 13, 19, 25, 31, 37, 43, 61, 67, 73, 79, 109, 139, 223 3 7 7, 8 4 7, 8, 29, 43 3 7, 8, 29, 43, 71, 113, 127 2 8 2, 9, 49 9 2, 9, 17, 49 5 2, 5, 9, 17, 41, 49 4 2, 5, 9, 13, 17, 29, 41, 49, 73, 89, 97, 113, 121, 137, 233, 257, 337, 761 3 9 3, 4 9 3, 4, 19 5 3, 4, 19, 37 3 3, 4, 7, 19, 37, 73, 109, 127, 163, 181, 199, 271, 307 2 10 2, 5, 81 11 2, 5, 11, 81 6 2, 5, 11, 31, 81 5 2, 5, 11, 31, 41, 61, 81 4 2, 5, 11, 31, 41, 61, 71, 81, 101, 131, 151, 181, 191, 211, 241, 251, 271, 281, 3 311, 331, 401, 421, 431, 461, 491, 641, 911 11 11, 23 5 11, 23, 67 4 11, 23, 67, 89 3 11, 23, 67, 89, 199, 331, 353, 419, 463, 617 2

Proof. Given a k value, we want to find an integer n with the property that every element of R can be written as a sum of at most n many kth powers in R. We want to use the previous theorem. One of the assumptions in that theorem is gcd( R , k) = 1. In the table, notice that q-column always has the divisors of k. So when | | we say q does not divide R in the statement, it is guaranteed that gcd( R , k) = 1. | | | |

If R is commutative, then R /J is isomorphic to the direct product of some finite fields as we discussed earlier. If R is not commutative, then let α R. Since Z[α] is a commutative finite ring with identity, we ∈ Z[α] have J (Z[α]) = Fq1 Fqr for some finite fields Fq1 , , Fqr . That means to find m in the ∼ ×···× · · · th previous theorem,. first we need to find how many k powers is necessary to write every element of Fqi as th a sum of k powers in Fqi . We denote this number with mi for each Fqi , i.e. every element of Fqi can be th R Z[α] written as a sum of mi many k powers in Fqi . Then, every element of /J or resp. J (Z[α]) th  .  can be written as a sum of m = max16i6r mi many k powers in R /J or resp. in Z[α] J (Z[α]) . Therefore, building on our results in Section 3 we can determine m, and using the previous theorem. we can let n = m or n = m + 1 depending on the parity of k.  30 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

Notice that in the previous theorem, we did not note the rings in which every element is a kth power, i.e. there are not any rows with n = 1 in the table. The following propositions are stated with this purpose, and their proofs follow from Proposition 3.1 and Theorem 5.1 easily.

Proposition 5.3. Let R be a finite ring with identity and with a cube-free order. Let R = pi1 pi2 pis be | | 1 2 · · · s the prime factorization of the order of the ring. Let k be a positive integer such that pj ∤ k for any j, and i gcd(k,p j 1) = 1 for all 1 6 j 6 s. If α R is a unit, then α is a kth power in R. If k is odd, then any j − ∈ element of the ring is a kth power. If k is even, then any element of the ring can be written as a sum of two kth powers.

i1 i2 is Proposition 5.4. Let R be a finite ring with identity. Let R = p1 p2 ps be the prime factorization of | | · · · i the order of the ring. Let k be a positive integer such that p ∤ k for any j, and gcd k, j (pι 1) = 1 j ι=1 j − for all 1 6 j 6 s. If α R is a unit, then α is a kth power in R. If k is odd, then any element of the ring is a ∈ Q kth power. If k is even, then any element of the ring can be written as a sum of two kth powers.

For example consider the group algebra F3[T ] where T is the group of upper triangular 3 3 matrices with 1 a b × 1’s in the diagonal over F i.e. T = 0 1 c a, b, c F . T is a group under multiplication and 3   ∈ 3 0 0 1   nonabelian, this implies F3[T ] is noncommutative.  Moreover, Maschke’s theorem does not apply, the group   algebra is not semisimple. But via the last proposition we can conclude that every element of F3[T ] is a septic (7th power), since 3 ∤ 7 and gcd(7, 2 8 26 80) = 1. × × ×

APPENDIX A.LEMMATA

Here we prove the basic inequalities used in Section 3.

Lemma A.1. Let x,y Z 1 . If y>x4, then (y 1)4 x4y3 + (y 1)y2x3 > 0. ∈ + \ { } − − −

Proof. Step 1: Pick any x Z 1 and fix it. ∈ + \ { } We will prove F (x,y) = (y 1)4 x4y3 + (y 1)y2x3 > 0 when y = x4 + 1. − − − F (x,x4 +1) = x16 x4(x4 + 1)3 + x4(x4 + 1)2x3 − = x4(x3 x 1)(x8 + x6 2x5 + 3x4 x3 + x2 x + 1). − − − − − x4 > 0 for every x R. ∈ x3 x 1 has only one real root in the interval (1, 2) and two complex conjugate roots, as the first derivative − − test and sign chart illustrates it. We have x3 x 1 > 0 when x Z 1 . − − ∈ + \ { } Also, x8 + x6 2x5 + 3x4 x3 + x2 x + 1 > 0 as x8 + x6 > 2x5 and 2x4 >x3 + x when x Z 1 . − − − ∈ + \ { } As a result, we proved F (x,y) > 0 when y = x4 + 1 and x Z 1 . ∈ + \ { } Step 2: Pick any x Z 1 . Now, we will prove by induction that F (x,y) > 0 for any y>x4. We ∈ + \ { } WARING’S PROBLEM IN FINITE RINGS 31 already showed in Step 1 that F (x,y) > 0 when y = x4 + 1. We assume this holds for y = x4 + C for any C Z , and we demonstrate below F (x,y) > 0 also holds when y = x4 + C + 1. ∈ + We have

F (x,x4 + C) = (x4 + C 1)4 x4(x4 + C)3 + (x4 + C 1)(x4 + C)2x3 > 0 − − − by assumption, and

F (x,x4 + C + 1) = (x4 + C + 1 1)4 x4(x4 + C + 1)3 + (x4 + C + 1 1)(x4 + C + 1)2x3. − − − Notice that if we show B = [F (x,x4 + C + 1) F (x,x4 + C)] > 0, then we are done. − We have

B = 4C3+9C2x4+3C2x3 6C2 +6Cx8+6Cx7 15Cx4 +Cx3+4C+x12+3x11 9x8 +x7+3x4 1 . − − − − Notice that we have | {z } | {z } | {z } | {z } 3C2x3 6C2 = 3C2(x3 2) > 0 − − 6Cx7 15Cx4 = 3Cx4(2x3 5) > 0 − − 3x11 9x8 = 3x8(x3 3) > 0 and 3x4 1 > 0 − − − since x Z 1 . These calculations show B > 0, and the result follows.  ∈ + \ { } Lemma A.2. Let x,y Z 1 . If y>x3, then (y 1)3 x2y(xy y + 1) > 0. ∈ + \ { } − − −

Proof. We again use induction to prove the claim. Step 1: Pick any x Z 1 and fix it. ∈ + \ { } We need to show F (x,y) = (y 1)3 x2y(xy y + 1) > 0 holds when y = x3 + 1. − − − We have F (x,x3 +1) = x8 2x6 + x5 x3 is clearly bigger than zero as x Z 1 . − − ∈ + \ { } Step 2: Pick any x Z 1 . We assume the claim holds for y = x3 + C for any C Z , and we ∈ + \ { } ∈ + demonstrate below F (x,y) > 0 holds also when y = x3 + C + 1. We have

F (x,x3 + C) = (x3 + C 1)3 x2(x3 + C)[x(x3 + C) (x3 + C) + 1] > 0 − − − by assumption, and

F (x,x3 + C + 1) = (x3 + C + 1 1)3 x2(x3 + C + 1)[x(x3 + C + 1) (x3 + C + 1) + 1]. − − − Notice that if we show B = [F (x,x3 + C + 1) F (x,x3 + C)] > 0, then we are done. − We have B = 3C2 + 4Cx3 + 2Cx2 3C +x6 + 2x5 4x3 +1. − − 2 5 3 Notice that both 2Cx 3C > 0 and 2x 4x > 0 as x Z+ 1 . So, we have B > 0 and the result − − | {z } ∈ \| { }{z } follows.  32 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

APPENDIX B.COMPUTER CODE

The computer search was carried out on a supercomputer using a computer program called Sage (ours was Sage/6.9). The code used to generate Table 1, 2, 3 and the data in Theorem 3.8 is provided below. Inside the code there are some parts written after the pound sign #. These are just comments for better understand- ing the code, and these comments are ignored by the computer. Also, note that we provided the original, indented code below. If you want to use the same code for some calculation, you need to preserve this indentation. k=37 def f(n): #We define a new function f(n) in the next four lines. A =[] #A is an empty list.

for j in FiniteField(n,′a′): #FiniteField(n,′a′) is the finite field with order n. A.append(j k) #This command adds jk to A. ∗ ∗ return uniq(A) #uniq(A) removes the same entries of the array A.

def f2(n) : A2=[] for x in f(n) : for y in f(n) : A2.append(x + y)

return uniq(A2) #f2(n) gives the list of the elements in 2Rk.

#Similarly, we have the definition of fX here for every X between 2 and 41. · · ·

def f41(n) : A41 = [ ] for x in f40(n) : for y in f(n) : A41.append(x + y)

return uniq(A41) #f41(n) gives the list of the elements in 41Rk. WARING’S PROBLEM IN FINITE RINGS 33

S =[] #S will keep the list of the finite fields with order n for n 6 8k4. for n in range(1, 8k4 + 1) : if len(prime_divisors(n))==1: #len(prime_divisors(n)) gives the number of distinct prime divisors of n. S.append(n)

S1=[] #S1 will keep the list of the finite fields with order n such that R = n. | k| 6 for n in S : if len(f(n))! = n : S1.append(n)

S2=[] #S2 will keep the list of the finite fields with order n such that 2R = n. | k| 6 for n in S1 : if len(f2(n))! = n : S2.append(n)

#Similarly, we have the definition of SX here for every X between 2 and 41. · · ·

S41=[] #S41 will keep the list of the finite fields with order n such that 41R = n. | k| 6 for n in S40 : if len(f41(n))! = n : S41.append(n)

print ′k =′, k print ′S1′,′ :′, S1 print ′S2′,′ :′, S2 #Similarly, we have the same command for SX here for every X between 2 and 41. · · · print ′S41′,′ :′, S41 34 YE¸SIM˙ DEMIRO˙ GLU˘ KARABULUT

ACKNOWLEDGEMENT

I would like to express my sincere gratitude to my advisers, Professor Jonathan Pakianathan and Professor David Covert for suggesting this problem. I also thank University of Rochester and the CIRC team for letting me use their supercomputer. This work was partially supported by NSA grant H98230-15-1-0319.

REFERENCES

[1] Alperin, Jonathan L. Local representation theory: Modular representations as an introduction to the local representation theory of finite groups. Vol. 11. Cambridge University Press, 1993. [2] Barbara, Roy. "91.07 Sums of cubes in finite fields." The Mathematical Gazette 91, no. 520 (2007): 85–87. [3] Bhaskaran, M. "Sums of m th powers in algebraic and abelian number fields." Archiv der Mathematik 17, no. 6 (1966): 497– 504. [4] Brouwer, Andries E., and Willem H. Haemers. Spectra of graphs. Springer Science & Business Media, 2011. [5] Curtis, Charles W., and Irving Reiner. Representation theory of finite groups and associative algebras. Vol. 356. American Mathematical Soc., 1966. [6] Farb, Benson, and R. Keith Dennis. Noncommutative algebra. Vol. 144. Springer Science & Business Media, 2012. [7] Godsil, Chris, and Gordon F. Royle. Algebraic graph theory. Vol. 207. Springer Science & Business Media, 2013. [8] Hungerford, Thomas W. Algebra. Vol. 73. Springer Science & Business Media, 2003. [9] Ireland, Kenneth, and Michael Rosen. A classical introduction to modern . Vol. 84. Springer Science & Business Media, 2013. [10] Katre, S., and Anuradha Garge. "Matrices over commutative rings as sums of k-th powers." Proceedings of the American Mathematical Society 141, no. 1 (2013): 103–113. [11] Robert, Alain M. A course in p-adic analysis. Vol. 198. Springer Science & Business Media, 2013. [12] Serre, Jean-Pierre. Linear representations of finite groups. Vol. 42. Springer Science & Business Media, 2012. [13] Small, Charles. "Solution of Waring’s Problem modn." The American Mathematical Monthly 84, no. 5 (1977): 356–359. [14] Small, Charles. "Sums of powers in large finite fields." Proceedings of the American Mathematical Society 65, no. 1 (1977): 35–36. [15] Small, Charles. "Waring’s problem." Mathematics Magazine 50, no. 1 (1977): 12–16. [16] Small, Charles. "Waring’s Problem modn." The American Mathematical Monthly 84, no. 1 (1977): 12–25. [17] Tornheim, Leonard. "Sums of n-th powers in fields of prime characteristic." Duke Math. J 4 (1938): 359–362. [18] Vaughan, Robert Ch, and Trevor D. Wooley. "Waring’s problem: a survey." Number theory for the millennium 3 (2002): 301–340. [19] Weil, André. "Numbers of solutions of equations in finite fields." Bull. Amer. Math. Soc 55, no. 5 (1949): 497–508. [20] West, Douglas Brent. Introduction to graph theory. Vol. 2. Upper Saddle River: Prentice hall, 2001. [21] Winterhof, Arne. "On Waring’s problem in finite fields." Acta Arithmetica 87, no. 2 (1998): 171–177.

E-mail address: [email protected]

DEPARTMENT OF MATHEMATICS,UNIVERSITY OF ROCHESTER,ROCHESTER, NY